Next Article in Journal
Mitigating Co Metal Particle Agglomeration and Enhancing ORR Catalytic Activity through Nitrogen-Enriched Porous Carbon Derived from Biomass
Previous Article in Journal
Facile Synthesis of Silicon-Based Materials Modified Using Zinc(Ⅱ) 2-Bromoacetic as Heterogeneous Catalyst for the Fixation of CO2 into Cyclic Carbonates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Palladium Nanoparticles Supported over Fused Graphene-like Material for Hydrogen Evolution Reaction

1
Applied Research Center at Thomas Jefferson National Accelerator Facility, Newport News, VA 23606, USA
2
Department of Molecular Biology and Chemistry, Christopher Newport University, Newport News, VA 23606, USA
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(7), 1117; https://doi.org/10.3390/catal13071117
Submission received: 20 June 2023 / Revised: 11 July 2023 / Accepted: 13 July 2023 / Published: 17 July 2023
(This article belongs to the Special Issue Novel Metal-Based Catalysts in Hydrogen Production)

Abstract

:
The search for a clean abundant energy source brought hydrogen gas into the limelight; however, the explosive nature of the gas brings up issues with its storage. A way to mitigate this danger is through the storing of hydrogen in a hydrogen feedstock material, which contains a large percentage of its weight as hydrogen. Sodium borohydride is a feedstock material that gained a lot of attention as it readily reacts with water to release hydrogen. This study explored a novel composite composed of palladium nanoparticles supported on a sugar-derived fused graphene-like material support (PdFGLM) for its ability to catalyze the reaction of sodium borohydride in water. Transmission electron microscopy (TEM), scanning electron microscopy (SEM), energy-dispersive X-ray spectroscopy (EDS), Fourier transform infrared spectroscopy (FTIR), and X-ray diffraction (XRD) were used to characterize and determine the size and shape of the catalyst used in this study. The XRD study detected the presence of palladium nanoparticles, and the EDS date confirmed the presence of 3% palladium nanoparticles. The TEM result shows the palladium nanoparticles of 5.5 nm incorporated to the graphene-like material layers. The composite contained approximately 3% palladium. In the hydrogenation reactions, it was observed that optimal reaction conditions included lower pHs, increased temperatures, and increased dosages of sodium borohydride. The reaction had the greatest hydrogen generation rate of 0.0392 mL min−1 mgcat−1 at pH 6. The catalyst was tested multiple times in succession and was discovered to increase the volume of hydrogen produced, with later trials indicating the catalyst becomes more activated with multiple uses. The activation energy of the reaction as catalyzed by PdFGLM was found to be 45.1 kJ mol−1, which is comparable to other catalysts for this reaction. This study indicates that this catalyst material has potential as a sustainable material for the generation of hydrogen.

1. Introduction

Fossil fuels currently dominate the world’s energy system; however, these fossil fuels are predicted to become depleted in the future [1,2]. The combustion of fossil fuel is responsible for a majority of anthropogenic emissions [2]. Moreover, fossil fuel combustion by-products posed a significant threat to children’s health [3]. They not only have significant impact on children’s physical development, but also greatly cause impairment in cognitive and behavioral development [3]. In order to address the issues, scientists are researching alternative eco-friendly energy sources [4,5]. Among the energy sources, hydrogen fuel arises as a potential candidate due to its power efficiency and cleaned combustion by-product. For example, the MARKAL model estimated that hydrogen can make up 22–50% of the electricity generation in Japan and reduce 80% of CO2 emissions [6]. Even though hydrogen energy has many advantages, it also has some disadvantages that prevent its large-scale application. For example, maximizing the compression of hydrogen into a storage system requires a large investment and has high operating costs due to the need to maintain the pressure and vessels [7,8]. In order to store more hydrogen, the popular method is to compress hydrogen into a gas cylinder under a pressure of 700 bar or higher; however, this method is often expensive and difficult to transport due to the weight of the gas cylinder [7,8]. Additionally, the temperature is another factor that affects the compression process. Hydrogen is under the reverse joule-Thomson effect, where hydrogen heats up when expanding at a temperature above (−80 °C) [7,8]. Another method of storing hydrogen is to convert it into liquid form. Hydrogen can be converted to liquid form at a temperature of 20–21 K; however, the liquefaction process is a time and energy consuming process. Furthermore, 40% energy of hydrogen is lost during the conversion process [7,8]. Hydrogen can be produced by the water electrolysis method; however, this method also consumes electrical energy [9,10].
Another cost-efficient and green way of hydrogen storage is using borohydride compounds [11]. Metal borohydrides are found to be a potential hydrogen storage material that can store at high density, but moderate temperature and hydrogen pressure [12]. Among the metal borohydride, sodium borohydride (NaBH4) is well-researched and appears to be the best hydrolytic compound for the broad scale generation of hydrogen [13]. The hydrolysis reaction of NaBH4 is shown in Equation (1):
NaBH4 + 2H2O → NaBO2 + 4H2.
The key challenge for applying NaBH4 to generate hydrogen is the slow reaction rate. There are many studies for optimal catalysts that can enhance the hydrogen evolution reaction [14,15,16]. Among catalysts, nanoparticles are well known for their ability to catalyze hydrogen generation reactions [17,18]. It was indicated in previous research that the size and availability of a support template can greatly affect the catalytic ability of nanoparticles [17,18,19]. Highly dispersed nanoparticles were found to exhibit a greater catalytic capability due to the high surface area-to-volume ratio; however, the nanoparticles tend to agglomerate or aggregate and form large nanoparticles [19,20]. The aggregation of nanoparticles leads to the loss of catalytic property [21]. It is necessary to utilize a support template to not only avoid the aggregation, but also improve the performance of nanoparticles [20,21,22,23,24]. Various templates were effectively applied in supporting nanoparticles such as graphene oxide, mesoporous silica, and carbon nanotubes; however, these materials are either expensive or require toxic chemicals or hazardous synthesis methods [25,26,27].
In our study, palladium nanoparticles are supported over fused graphene-like material (PdFGLM) and were applied as a catalyst for hydrogen evolution reaction of NaBH4. Palladium is in the noble metal group and was extensively researched for fuel cell development due to its catalytic and hydrogen absorbing properties [28,29,30]. Palladium is widely used as catalysts in various chemical reactions due to its unique properties. Palladium nanoparticles possess a high surface-to-volume ratio, which provides a large number of active sites for catalytic reactions [31,32,33]. The small size of the nanoparticles enhances their reactivity and promotes efficient catalytic activity. For example, palladium nanoparticles catalyze the addition of hydrogen to unsaturated organic compounds, such as alkenes and alkynes, leading to the formation of saturated compounds. Palladium nanoparticles are used in various cross-coupling reactions, such as Suzuki–Miyaura, Heck, and Sonogashira reactions [31,32,33]. These reactions involve the formation of carbon–carbon or carbon–heteroatom bonds. Palladium nanoparticles can catalyze the oxidation of organic compounds using molecular oxygen as the oxidant. This includes the selective oxidation of alcohols to aldehydes or ketones. Palladium nanoparticles are employed in the synthesis of complex organic molecules, such as natural products and pharmaceuticals, by facilitating carbon–carbon bond formation reactions such as Heck and Stille reactions [31,32,33]. Palladium nanoparticles are utilized as catalysts in fuel cells to promote the oxidation of hydrogen or methanol, generating electricity. In previous studies, palladium nanoparticles were shown to effectively enhance the hydrogen evolution reaction [28,29,30,31,32,33]. Palladium nanoparticles can be synthesized using various methods, including chemical reduction, thermal decomposition, and colloidal synthesis. Control over the size, shape, and surface chemistry of the nanoparticles allows for tailoring their catalytic properties. Overall, palladium nanoparticles are versatile catalysts widely used in organic synthesis, pharmaceutical manufacturing, fine chemical production, and environmental applications due to their excellent catalytic properties [28,29,30,31,32,33]. Their small size and large surface area contribute to efficient catalytic activity, reducing reaction times and increasing reaction yields [28,29,30,31,32,33].
The fused graphene-like material was used in this study as a support template for enhancing the dispersing of palladium nanoparticles. Since the discovery of graphene material, there was much research focusing on materials with similar structure and property to that of graphene [34,35]. The synthesis of graphene often requires toxic chemicals and poses a high risk of combustion and explosion accident [36,37,38]. Our fused graphene-like materials (FGLM) can be safely synthesized from biomass such as glucose and plant-based material. Previous researchers showed the effect of FGLM in controlling the size and distribution of palladium nanoparticles [39,40]. For example, in a study by Klyuev et al., the graphene sheet succeeds in enhancing the uniform distribution of small-sized Pd nanoparticles [40].
Graphene-like materials refer to a class of materials that possess similar structural and electronic properties to graphene [41]. While graphene is a single layer of carbon atoms arranged in a two-dimensional honeycomb lattice, graphene-like materials can include other two-dimensional materials that exhibit similar characteristics. For example, graphene is a prototypical graphene-like material. It consists of a single layer of carbon atoms arranged in a hexagonal lattice [41]. Graphene exhibits exceptional electrical conductivity, mechanical strength, and thermal conductivity, making it a highly sought-after material for various applications. Additionally, graphene oxide is derived from graphene by introducing oxygen-containing functional groups onto the graphene lattice. This modification imparts hydrophilicity to the material and makes it suitable for applications such as water purification, energy storage, and biomedicine. In addition, graphene nanoribbons are narrow strips of graphene with a width in the nanometer range. They possess unique electronic properties, such as bandgap opening, which makes them potential candidates for electronic devices, including transistors and integrated circuits. Graphene with 3–4 layers of carbon is commonly referred to as few-layer graphene or graphene-like material. It is a stacked arrangement of graphene sheets, with each sheet consisting of a single layer of carbon atoms arranged in a hexagonal lattice. Few-layer graphene can be synthesized through several methods, including mechanical exfoliation, chemical vapor deposition (CVD), and epitaxial growth [41]. The specific synthesis technique can influence the quality, layer thickness, and uniformity of the few-layer graphene produced. Few-layer graphene exhibits many of the properties of single-layer graphene, but some properties may be influenced by the layer stacking. Key properties of few-layer graphene include high electrical conductivity, excellent mechanical strength, exceptional thermal conductivity, and a large surface area [41].
In our work, the synthesized PdFGLM composite was characterized by powdered X-ray diffraction (P-XRD), scanning electron microscopy–energy dispersive spectroscopy (SEM-EDS), Fourier transform infrared spectroscopy (FTIR), and transmission electron microscopy (TEM). The catalytic ability of PdFGLM was verified in the hydrogen generation reaction under various conditions (pH, temperature, and dosage of reactants). The durability of the composite was examined through reusability trials.

2. Results

2.1. Characterization

TEM imaging was used to characterize the novel PdFGLM catalyst as shown in Figure 1. At a scale of 100 nm (Figure 1A), the FGLM backbone can be observed with nanoparticles dispersed across its surface. Similar to the other composites, the FGLM appears to be somewhat round masses fused together; however, this material seems to be more fused into a larger mass than the others. Figure 1C allowed for the average diameter of the PdNPs to be determined to be about 5.5 nm. The interplanar spacings of PdNPs were depicted in Figure 1D. The d-spacing was estimated to be 0.12 nm.
The presence of palladium metal in the novel PdFGLM catalyst was confirmed via EDS analysis (Figure 2). Only carbon, oxygen, and palladium were seen in the sample at weight percentages of 48%, 49%, and 3%, respectively. From Figure 1, it was determined that the PdNPs were very small, which explains the low weight percentage of palladium compared to the other elements.
The PdFGLM catalyst was characterized using XRD analysis and compared to FGLM material containing no nanoparticles (Figure 3). There is a clear broad peak seen in both the unsupported FGLM material and the composite around 22 degrees. This peak is commonly seen in carbon-based materials, as it is representative of graphitic characteristics [42]. Many peaks below 40° were attributed to the structure of beta cyclodextrin. Those peaks were consistent with the peaks of beta cyclodextrin reported in the Song et al. study of 2011 [43]. For example, the peaks at 6°, 9°, 11°, 12°, 15°, 17°, 19°, 21°, 23°, and 27° were attributed to the (001), (101), (130), (041), (141), (180), (230), (042), (162), and (222) planes of beta cyclodextrin, respectively. The other peaks at 40°, 46°, 68°, and 82° corresponded to the (111), (200), (220), and (311) lattice planes of palladium nanoparticles (JCPDS 05-0681). The combination of TEM and EDS seen in Figure 1 and Figure 2 confirmed the presence of incredibly small palladium nanoparticles.
The PdFGLM catalyst was then characterized using FTIR analysis as shown in Figure 4. When looking at the unsupported FGLM material, a broad peak that is characteristic of -OH functional groups can be seen ranging from ~3600 cm−1 to ~3000 cm−1. The chemical structure of dextrose, which was used to synthesize the FGLM material, contains multiple -OH groups, which is likely the cause for this stretch. A small peak at ~2900 cm−1 is indicative of the C-C bond of the dextrose. Another peak was seen at ~1700 cm−1 that can be attributed to the C=O of the dextrose molecule. All three of these groups can also be seen in the composite material with a slight shift of the peaks due to the presence of PdNPs.

2.2. Catalytic Evaluation with Varied Dosage of Reactants

The first catalytic condition the PdFGLM composite was tested for was under varying dosages of NaBH4 (Figure 5). The catalyst was first tested at 925 μmol, which would become the standard dose for comparison. At this dosage, the reaction produced 23.2 mL of hydrogen over the course of two hours, resulting in a production rate of 0.0196 mL min−1 mgcat−1. The dose of NaBH4 was then lowered to 625 μmol, which resulted in 19.6 mL of hydrogen produced at a rate of 0.0141 mL min−1 mgcat−1. Lastly, when the dose was raised to 1225 μmol, a rate of 0.0241 mL min−1 mgcat−1 was observed, producing 28.9 mL of hydrogen. There is a clear trend that can be observed through Figure 5, where increased doses of NaBH4 resulted in increased volumes of hydrogen being produced. These results are supported by the previous work [14,44]. Based on theoretical calculation, each one mole of NaBH4 can produce 89.6 L of hydrogen gas at standard pressure and temperature (STP).

2.3. Catalytic Evaluation at Varied pH Conditions

The next condition PdFGLM was tested at was under varying pH conditions (Figure 6). The first pH test was pH 7, which would then be used as a standard for comparison. At pH 7, the reaction had a generation rate of 0.0196 mL min−1 mgcat−1 and produced roughly 23.2 mL after 2 h. Next, the pH of the solution was lowered to 6, which produced 47.0 mL after two hours at a rate of 0.0392 mL min−1 mgcat−1. Lastly, when the pH of the reaction was raised to 8, the hydrogen generation rate was observed to be 0.0186 mL min−1 mgcat−1, producing 22.3 mL after 2 h. These results indicate that this reaction as catalyzed by PdFGLM produces more hydrogen at lower pHs and less at higher pHs. Both those results are supported by the previous work of Schlesinger et al. (1953) and Kaufman et al. (1985), respectively [14,44]. At pH 7, it was theoretically calculated that each one mole of NaBH4 can produce 89,600 mL of H2 at STP.

2.4. Catalytic Evaluation at Varied Temperatures

The catalytic ability of PdFGLM was also tested under different temperatures as seen in Figure 7. The first temperature tested was room temperature (RT 295 K), which would be the standard for comparison. At RT, the volume of hydrogen produced was found to be 23.2 mL after two hours at a generation rate of 0.0196 mL min−1 mgcat−1. The temperature was then lowered to 288 K, which resulted in a hydrogen generation rate of 0.0098 mL min−1 mgcat−1 and a volume of 11.7 mL. When the temperature was lowered even further to 283 K, the volume of hydrogen produced was found to be 8.9 mL with a generation rate of 0.0074 mL min−1 mgcat−1. The temperature was then raised to 303 K, which resulted in a hydrogen generation rate of 0.0244 mL min−1 mgcat−1 and a volume of 29.3 mL. These results indicate that there is a direct relationship between temperature and the volume of hydrogen produced. At 295 k, the theoretical volume of H2 produced per mole of NaBH4 was 89,600 mL at STP.

2.5. Activation Energies

With the completion of the temperature study, an Arrhenius plot (Figure 8) was created using each temperature tested and the Arrhenius Equation (2). Based on the Arrhenius plot, it indicated that the reaction as catalyzed by PdFGLM is endothermic.
k = A e E a R T
R: universal gas constant;
T: Temperature in K;
Ea: Activation energy;
A: Pre-exponential factor;
k: rate constant.
Figure 8. Arrhenius plot used to determine the activation energy of the NaBH4 reaction catalyzed by PdFGLM.
Figure 8. Arrhenius plot used to determine the activation energy of the NaBH4 reaction catalyzed by PdFGLM.
Catalysts 13 01117 g008
Once Figure 8 was created, the equation of the line could be determined. Based on the inverse of the slope of that line multiplied by the universal gas constant, the activation energy of this reaction as catalyzed by PdFGLM was determined to be 45.1 kJ/mol. This activation energy was then compared to other similar catalysts for the hydrolysis of NaBH4 in Table 1. When compared to non-precious metals such as nickel and cobalt, PdFGLM holds a clear improvement in the activation energy of the reaction. This catalyst also has a lower activation energy than other precious metal composite catalysts using metals such as ruthenium and gold, and is only slightly higher than a silver composite. On palladium-based composite, Pd/C (commercial catalyst) had a lower activation energy than PdFGLM, yet Pd/MWCNTs had a higher activation energy, indicating that the type of support material chosen to make the composite catalyst plays a role in the catalytic ability of the metal. Despite not having the lowest reported activation energy for this reaction, PdFGLM does show competitiveness, performing as well if not better than many similar catalysts. Additionally, many of these other composites require the use of high temperatures or harsh chemicals, whereas PdFGLM uses a relatively mild temperature and only sugar solution to be made.

2.6. Catalytic Reusability of Palladium Composite

The final catalytic study conducted on PdFGLM was its reusability (Figure 9). The same amount of catalyst was tested for five consecutive trials. The initial trial was run at 295 K, pH 7, and used a dose of 925 μmol of NaBH4. This trial ran for approximately two hours, at which point an additional 925 μmol of NaBH4 was added, marking the start of the second two-hour trial. This was carried out for a total of five trials. Based on the data, it appears that the catalyst remains stable throughout the five trials, with no loss in catalytic ability and an average volume of 30.8 mL across the trials. The first three trails produced a consistent amount of hydrogen, with the latter two increasing the volume. The hydrogen generation over time of each trial is shown in Supplementary Material (Figure S1). Theoretically, it was calculated that each 925 μmol of sodium borohydride produced 90 mL of hydrogen.
Equation (3) showed the binding of BH4 and H species on the surface of PdNPs, which can explain the increasing in the hydrogen generated in the reusability study. Equation (4) shows that the bonds become hydrolyzed, and this process improves the electrostatic stabilization of the PdNPs surface, making them more active [59].
2   Surface-Pd + BH 4     Surface-Pd-H + Surface-Pd- BH 3
Surface-Pd-BH3 + Surface-Pd-H + HOH → Surface-Pd-[BH2(OH)] + H2 + Surface-Pd
This could indicate that the catalyst is becoming more activated with each use. This apparent stability and possible activation make this PdFGLM catalyst an even more attractive option for the hydrolysis of NaBH4. Additionally, after the reusability trials, the PdFLGM was re-collected and characterized by P-XRD and FTIR. The P-XRD spectra (Figure S2) showed that the material still had a graphitic peak at 20° and all PdNPs peaks at 40°, 46°, 68°, and 82°. However, the beta-cyclodextrin peaks were no longer presented in the P-XRD. It highly indicated that the capping agent was removed during the hydrolysis trials. Due to the loss of capping agents, it created more space for BH4 to surround and stabilize the PdNPs. Through the FTIR spectra of sodium borohydride and PdGLM after reusability (Figures S3 and S4), it depicted that the B-H bending group and B-H stretching groups presented on the PdFGLM materials after reusability trials.

3. Proposed Mechanism

This team assumed that the mechanism by which PdFGLM catalyzes the reaction of NaBH4 in water occurs as shown in Scheme 1. The synthesis of the composite material resulted in palladium nanoparticles affixed to the surface of the FGLM material. In solution, a borohydride ion (BH4) bonds to one of the PdNPs. A nearby water molecule will then bond to the boron, with one hydrogen bonding to a hydrogen from the borohydride and being released as a H2 gas molecule. This will happen three more times, at which point the resulting tetrahydroxyborate ion [B(OH4)] separates from the PdNP and another BH4 molecule takes its place. The mechanism was similar to the study of Guella et al. (2006) [60].

4. Experimental

4.1. Synthesis of Fused Graphene-like Material (FGLM)

The synthesis of fused graphene-like material (FGLM) followed the method of Adel et al. with a few modification [37]; 0.5 M of dextrose solution was prepared and added to 50 mL Teflon-lined autoclave tubes at a 3:2 ratio of vapor to liquid. The autoclave tubes were heated at 200 °C for 4 h and then cooled to room temperature. The product was collected and centrifuged at 15,000 rpm for 30 min. The collect precipitate was washed with DI water several times and then dried in the oven.

4.2. Synthesis of Palladium Nanoparticle (PdNPs) and PdFGLM

The palladium nanoparticles (PdNPs) were synthesized by following the method of Wen et.al. [61], and 1 mM of palladium chloride solution was prepared and mixed with 48 mL of 10 mM beta cyclodextrin solution. The mixture was stirred for 10 min. After that, 0.25 mL of 180 mM NaBH4 solution was added to the mixture and stirred for 2 h. The nanoparticle solution was centrifuged at 10,000 rpm for 15 min to remove unreacted reactants.
For the synthesis of palladium nanoparticles supported over fused graphene-like material (PdFGLM). The nanoparticles were embedded on the FGLM by the incipient wetness method. A few milliliters of nanoparticle solution were added to the FGLM and then dried in the oven at 50 °C. The process was repeated several times.

4.3. Characterization

The crystallinity of FGLM and PdNPs was confirmed through powder X-ray diffraction (PXRD, Rigaku Miniflex II, Cu Kα X-ray, and nickel filters, Tokyo, Japan). The functional group of the nanocomposite was identified by Fourier transform infrared spectroscopy (FTIR, Shimadzu IR-Tracer 100, Kyoto, Japan) with an attenuated total reflectance (ATR, Shimadzu QATR-S, Kyoto, Japan) attachment. The ratio of PdNPs to FGLM was determined by scanning electron microscopy (SEM, JEOL JSM-6060LV, Tokyo, Japan) and energy dispersive X-ray spectroscopy (EDS, ThermoScientific UltraDry, Waltham, MA, USA).
In the preparation of the sample for transmission electron microscopy (TEM, JEM-2100F, Tokyo, Japan), the nanocomposite powder was mixed with DI water and sonicated for 10 min. A few drops of the solution were added to the TEM grid and dried in the oven.

4.4. Catalysis

The gravimetric water displacement system was set up to monitor the volume of generated hydrogen [15]. One Buchner flask was filled with 100 mL of DI water and was mixed with NaBH4 and 0.1 g of PdFGLM. The flask was closed and connected with another flask that contained 100 mL DI water through plastic tubing. The water displacement was monitored by the data logging software (SPDC Data Collector, OHAUS). The catalyzed reactions were examined under various pH (pH 6, pH 7, and pH 8), temperatures (283 K, 288 K, 295 K, and 303 K), and dosages of NaBH4 (625 μmol, 925 μmol, and 1225 μmol). Each reaction was monitored for approximately 2 h. The pH of solvent was adjusted by nitric acid and sodium hydroxide. In the reusability trials, the reaction occurred at pH 7 and at 295 K. The amount of PdFGLM remained the same, and 925 μmol NaBH4 was re-supplied after each reusability trial.

5. Conclusions

In conclusion, we report a novel sugar-derived palladium nanoparticle composite for the catalysis of NaBH4. This catalyst was characterized via TEM, SEM, EDS, XRD, and FTIR, confirming the identity of elements and functional groups within the sample. The catalyst was tested under various reaction conditions and it was observed that optimal reaction conditions include higher temperatures, lower pHs, and higher doses of NaBH4. This reaction produced the most hydrogen at pH 6, with a production rate of 0.0392 mL min−1 mgcat−1. It was found that the catalyst was able to be used multiple times in succession, producing an average of 30.8 mL across five trials, with the latter two trials producing increasing volumes of hydrogen, indicating the activation of the catalyst. The activation energy of the catalyst was found to be 45.1 kJ mol−1, which is comparable to other catalysts for this reaction. This comparable Ea, along with the reusability of the material and the sugar-based support, makes this a potentially sustainable option for the generation of hydrogen gas for use as a fuel source.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13071117/s1, Figure S1: Catalytic reusability of PdFGLM after 5 consecutive trials.; Figure S2: P-XRD of PdFGLM after reusability trials. Figure S3: FTIR of sodium borohydride (NaBH4). Figure S4: FTIR of PdFGLM after reusability trials.

Author Contributions

Q.Q.: data curation, formal analysis, writing—original draft. E.B.: data curation, formal analysis, writing—original draft. T.M.A.-F.: conceptualization, validation, formal analysis, investigation, resources, supervision, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

Corresponding Author acknowledges Lawrence J. Sacks Professorship in Chemistry.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hook, M.; Li, J.; Johansson, K.; Snowden, S. Growth Rates of Global Energy Systems and Future Outlooks. Nat. Resour. Res. 2011, 21, 23–41. [Google Scholar] [CrossRef]
  2. Hook, M.; Tang, X. Depletion of fossil fuels and anthropogenic climate change—A review. Energy Policy 2013, 52, 797–809. [Google Scholar] [CrossRef] [Green Version]
  3. Perera, F. Pollution from Fossil-Fuel Combustion is the Leading Environmental Threat to Global Pediatric Health and Equity: Solutions Exist. Int. J. Environ. Res. Public Health 2018, 15, 16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Brandon, N.P.; Kurban, Z. Clean energy and the hydrogen economy. Phil. Trans. R. Soc. A 2017, 375, 20160400. [Google Scholar] [CrossRef] [Green Version]
  5. Bosnjakovic, M.; Sinaga, N. The Perspective of Large-Scale Production of Algae Biodiese. Appl. Sci. 2020, 10, 8181. [Google Scholar] [CrossRef]
  6. Ozawa, A.; Kudoh, Y.; Murata, A.; Honda, T.; Saita, I.; Takagi, H. Hydrogen in low-carbon energy systems in Japan by 2050: The uncertainties of technology development and implementation. Int. J. Hydrogen Energy 2018, 43, 18083–18094. [Google Scholar] [CrossRef]
  7. Hua, T.Q.; Ahluwalia, R.K. Alane hydrogen storage for automotive fuel cells—Off-board regeneration processes and efficiencies. Int. J. Hydrogen Energy 2011, 36, 15259–15265. [Google Scholar] [CrossRef]
  8. Witkowski, A.; Rusin, A.; Majkut, M.; Stolecka, K. Comprehensive analysis of hydrogen compression and pipeline transportation from thermodynamics and safety aspects. Energy 2017, 141, 2508–2518. [Google Scholar] [CrossRef]
  9. Zhang, L.; Li, L.; Liang, J.; Fan, X.; He, X.; Chen, J.; Li, J.; Li, Z.; Cai, Z.; Sun, S.; et al. Highly efficient and stable oxygen evolution from seawater enabled by a hierarchical NiMoSx microcolumn@NiFe-layered double hydroxide nanosheet array. Inorg. Chem. Front. 2023, 10, 2766–2775. [Google Scholar] [CrossRef]
  10. Li, L.; Zhang, L.; Gou, L.; Wei, S.; Hou, X.; Wu, L. High-performance methanol electrolysis towards energy-saving hydrogen production: Using Cu2O-Cu decorated Ni2P nanoarray as bifunctional monolithic catalyst. Chem. Eng. J. 2023, 454, 140292. [Google Scholar] [CrossRef]
  11. Puszkiel, J.; Garroni, S.; Milanese, C.; Gennari, F.; Klassen, T.; Dornheim, M.; Pistidda, C. Tetrahydroborates: Development and Potential as Hydrogen Storage Medium. Inorganics 2017, 54, 74. [Google Scholar] [CrossRef] [Green Version]
  12. Li, H.W.; Yan, Y.; Orimo, S.I.; Zuttel, A.; Jesen, C.M. Recent progress in metal borohydrides for hydrogen storage. Energies 2011, 4, 185–214. [Google Scholar] [CrossRef]
  13. Demirci, U.B. About the Technological Readiness of the H2 Generationby Hydrolysis of B(-N)-H Compounds. Energy Technol. 2018, 6, 470–486. [Google Scholar] [CrossRef] [Green Version]
  14. Schlesinger, H.I.; Brown, H.C.; Finholt, A.E.; Gilbreath, J.R.; Hoekstra, H.R.; Hyde, E.K. Sodium borohydride, its hydrolysis and its use as a reducing agent and in the generation of hydrogen1. J. Am. Chem. Soc. 1953, 75, 215–219. [Google Scholar] [CrossRef]
  15. Dushatinski, T.; Huff, C.; Abdel-Fattah, T.M. Characterization of electrochemically deposited films from aqueous and ionic liquid cobalt precursors toward hydrogen evolution reactions. Appl. Surf. Sci. 2016, 385, 282–288. [Google Scholar] [CrossRef]
  16. Huff, C.; Long, J.M.; Heyman, A.; Abdel-Fattah, T.M. Palladium Nanoparticle Multiwalled Carbon Nanotube Composite as Catalyst for Hydrogen Production by the Hydrolysis of Sodium Borohydride. ACS Appl. Energy Mater. 2018, 1, 4635–4640. [Google Scholar] [CrossRef]
  17. Huff, C.; Dushatinski, T.; Barzanji, A.; Abdel-Fattah, N.; Barzanji, K.; Abdel-Fattah, T.M. Pretreatment of Gold Nanoparticle Multi-Walled Carbon Nanotube Composites for Catalytic Activity toward Hydrogen Generation Reaction. ECS J. Solid State Sci. Technol. 2017, 6, M69–M71. [Google Scholar] [CrossRef]
  18. Huff, C.; Long, J.M.; Abdel-Fattah, T.M. Beta-Cyclodextrin Assisted Synthesis of Silver Nanoparticle Network and its Application in a Hydrogen Generation Reaction. Catalysts 2020, 10, 1014. [Google Scholar] [CrossRef]
  19. Osborne, J.; Horten, M.; Abdel-Fattah, T.M. Gold Nanoparticles Supported Over Low-Cost Supports for Hydrogen Generation from a Hydrogen Feedstock Material. ECS J. Solid State Sci. Technol. 2020, 9, 071004. [Google Scholar] [CrossRef]
  20. Kushch, S.D.; Kujunko, N.S.; Tarasov, B.P. Platinum nanoparticles on carbon nanomaterials with graphene structure as hydrogenation catalysts. Russ. J. Gen. Chem. 2009, 79, 706–710. [Google Scholar] [CrossRef]
  21. Hitrik, M.; Sasson, Y. Aggregation of catalytically active Ru nanoparticles to inactive bulk, monitored in situ during an allylic isomerization reaction. Influence of solvent, surfactant and stirring. RSC Adv. 2018, 8, 1481–1492. [Google Scholar] [CrossRef] [PubMed]
  22. Tang, Z.; Yin, X.; Zhang, Y.; Zhang, N.; Xu, Y. Inhibiting Pd nanoparticle aggregation and improving catalytic performance using one-dimensional CeO2 nanotubes as support. Chin. J. Catal. 2013, 34, 1123–1127. [Google Scholar] [CrossRef]
  23. Quach, Q.; Abdel-Fattah, T.M. Silver Nanoparticles functionalized Nanoporous Silica Nanoparticle grown over Graphene Oxide for enhancing Antibacterial effect. Nanomaterials 2022, 12, 3341. [Google Scholar] [CrossRef] [PubMed]
  24. Biehler, E.; Whiteman, R.; Lin, P.; Zhang, K.; Baumgart, H.; Abdel-Fattah, T.M. Controlled Synthesis of ZnO Nanorods Using Different Seed Layers. ECS J. Solid State Sci. Technol. 2020, 9, 121008. [Google Scholar] [CrossRef]
  25. Rhazouani, A.; Gamrani, H.; Achaby, M.E.; Aziz, K.; Gebrati, L.; Uddin, M.S.; Aziz, F. Synthesis and Toxicity of Graphene Oxide Nanoparticles: A Literature Review of In Vitro and In Vivo Studies. Biomed. Res. Int. 2021, 2021, 5518999. [Google Scholar] [CrossRef]
  26. Mahmoud, A.M.; Desouky, E.M.; Hozayen, W.G.; Bin-Jumah, M.; El-Nahass, E.S.; Soliman, H.A.; Farghali, A.A. Mesoporous Silica Nanoparticles Trigger Lier and Kidney Injury and Fibrosis Via Altering TLR4/NF-κB, JAK2/STAT3 and Nrf2/HO-1 Signaling in Rats. Biomolecules 2019, 9, 528. [Google Scholar] [CrossRef] [Green Version]
  27. Zhang, C.; Wu, L.; Perror, M.D.; Zhao, X. Carbon Nanotubes: A Summary of Beneficial and Dangerous Aspects of an Increasingly Popular Group of Nanomaterials. Front. Oncol. 2021, 11, 693814. [Google Scholar] [CrossRef]
  28. Antolini, E. Palladium in fuelcell catalysis. Energy Environ. Sci. 2009, 2, 915–931. [Google Scholar] [CrossRef]
  29. Adams, B.D.; Chen, A. The role of palladium in a hydrogen economy. Mater. Today 2011, 14, 282–289. [Google Scholar] [CrossRef]
  30. Mahesh, K.N.; Balaji, R.; Dhathathreyan, K.S. Palladium nanoparticles as hydrogen evolution reaction (HER) electrocatalyst in electrochemical methanol reformer. Int. J. Hydrogen Energy 2016, 41, 46–51. [Google Scholar] [CrossRef]
  31. Albano, G.; Petri, A.; Aronica, A.A. Palladium Supported on Bioinspired Materials as Catalysts for C–C Coupling Reactions. Catalysts 2023, 13, 210. [Google Scholar] [CrossRef]
  32. Kathiresan, M. Metal-Ion/Metal Nanoparticle-Anchored Porous Organic Polymers as Efficient Catalysts for Organic Transformations—A Recent Overview. Chem. Asian J. 2023, 18, e202201299. [Google Scholar] [CrossRef]
  33. Cao, Y.; Ran, R.; Wu, X.; Si, Z.; Kang, F.; Weng, D. Progress on metal-support interactions in Pd-based catalysts for automobile emission control. J. Environ. Sci. 2023, 125, 401–426. [Google Scholar] [CrossRef]
  34. Karthick, K.; Bijoy, T.K.; Sivakumaran, A.; Basha, A.B.M.; Murugan, P.; Kundu, S. Enhancing Hydrogen Evolution Reaction Activities of 2H-Phase VS2 Layers with Palladium Nanoparticles. Inorg. Chem. 2020, 59, 10197–10207. [Google Scholar] [CrossRef]
  35. Tang, Q.; Zhou, Z.; Chen, Z. Innovation and discovery of graphene-like materials via density-functional theory computations. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2015, 5, 360–379. [Google Scholar] [CrossRef]
  36. Das, V.K.; Shifrina, Z.B.; Bronstein, L.M. Graphene and graphene-like materials in biomass conversion: Paving the way to the future. J. Mater. Chem. A 2017, 5, 25131–25143. [Google Scholar] [CrossRef] [Green Version]
  37. Adel, M.; El-Maghraby, A. Synthesis of few-layer graphene-like nanosheets from glucose: New facile approach for graphene-like nanosheets large-scale production. J. Mater. Res. 2016, 31, 455–467. [Google Scholar] [CrossRef]
  38. Zhou, Y.; Xie, C.; Su, L.; Han, X. Study of the risks of the graphene oxide preparation process by reaction calorimetry. J. Therm. Anal. Calorim. 2020, 139, 101–112. [Google Scholar] [CrossRef]
  39. Chen, Z.; Wang, W.; Zhang, Y.; Liang, Y.; Cui, Z.; Wang, X. Pd Nanoparticles Confined in the Porous Graphene-like Carbon Nanosheets for Olefin Hydrogenation. Langmuir 2018, 34, 12809–12814. [Google Scholar] [CrossRef]
  40. Klyuev, M.V.; Arbuzov, A.A.; Magdalinova, N.A.; Kalmykov, P.A.; Tarsov, B.P. Palladium-containing graphene-like material: Synthesis and catalytic activity. Russ. J. Phys. Chem. A 2016, 90, 1749–1753. [Google Scholar] [CrossRef]
  41. Xu, M.; Liang, T.; Shi, M.; Chen, H. Graphene-Like Two-Dimensional Materials. Chem. Rev. 2013, 113, 3766–3798. [Google Scholar] [CrossRef] [PubMed]
  42. Li, K.; Liu, Q.; Cheng, H.; Hu, M.; Zhang, S. Classification and Carbon Structural Transformation from Anthracite to Natural Coaly Graphite by XRD, Raman spectroscopy, and HRTEM. Spectrochim. Acta Part A 2021, 249, 119286. [Google Scholar] [CrossRef] [PubMed]
  43. Song, W.; Yu, X.; Wang, S.; Blasier, R.; Markel, D.C.; Mao, G.; Shi, T.; Ren, W. Cyclodextrin-erythromycin complexes as a drug delivery device for orthopedic application. Int. J. Nanomed. 2011, 6, 3173–3186. [Google Scholar] [CrossRef] [Green Version]
  44. Kaufman, C.M.; Sen, B. Hydrogen generation by hydrolysis of sodium tetrahydroborate: Effects of acids and transition metals and their salts. J. Chem. Soc. Dalton Trans. 1985, 2, 307–313. [Google Scholar] [CrossRef]
  45. Zhang, J.S.; Delgass, W.N.; Fisher, T.S.; Gore, J.P. Kinetics of Ru-catalyzed sodium borohydride hydrolysis. J. Power Sources 2007, 164, 772–781. [Google Scholar] [CrossRef]
  46. Liang, Y.; Dai, H.-B.; Ma, L.-P.; Wang, P.; Cheng, H.-M. Hydrogen generation from sodium borohydride solution using a ruthenium supported graphite catalyst. Int. J. Hydrogen Energy 2010, 35, 3023–3028. [Google Scholar] [CrossRef]
  47. Patel, N.; Patton, B.; Zanchetta, C.; Fernandes, R.; Guella, G.; Kale, A.; Miotello, A. Pd-C power and thin film catalysts for hydrogen production by hydrolysis of sodium borohydride. Int. J. Hydrogen Energy 2008, 33, 287–292. [Google Scholar] [CrossRef]
  48. Peña-Alonso, R.; Sicurelli, A.; Callone, E.; Carturan, G.; Raj, R. A picoscale catalyst for hydrogen generation from NaBH4 for fuel cells. J. Power Sources 2007, 165, 315–323. [Google Scholar] [CrossRef]
  49. Huff, C.; Dushatinski, T.; Abdel-Fattah, T.M. Gold Nanoparticle/Multi-Walled Carbon Nanotube Composite as Novel Catalyst for Hydrogen Evolution Reactions. Int. J. Hydrogen Energy 2017, 42, 18985–18990. [Google Scholar] [CrossRef]
  50. Huff, C.; Long, J.M.; Aboulatta, A.; Heyman, A.; Abdel-Fattah, T.M. Silver Nanoparticle/Multi-Walled Carbon Nanotube Composite as Catalyst for Hydrogen Production. ECS J. Solid State Sci. Technol. 2017, 6, M115–M118. [Google Scholar] [CrossRef]
  51. Huff, C.; Quach, Q.; Long, J.M.; Abdel-Fattah, T.M. Nanocomposite Catalyst Derived from Ultrafine Platinum Nanoparticles and Carbon Nanotubes for Hydrogen Generation. ECS J. Solid State Sci. Technol. 2020, 9, 101008. [Google Scholar] [CrossRef]
  52. Quach, Q.; Biehler, E.; Elzamzami, A.; Huff, C.; Long, J.M.; Abdel-Fattah, T.M. Catalytic Activity of Beta-Cyclodextrin-Gold Nanoparticles Network in Hydrogen Evolution Reaction. Catalysts 2021, 11, 118. [Google Scholar] [CrossRef]
  53. Huff, C.; Biehler, E.; Quach, Q.; Long, J.M.; Abdel-Fattah, T.M. Synthesis of Highly Dispersive Platinum Nanoparticles and their Application in a Hydrogen Generation Reaction. Colloids Surf. A. 2021, 610, 125734. [Google Scholar] [CrossRef]
  54. Biehler, E.; Quach, Q.; Huff, C.; Abdel-Fattah, T.M. Organo-Nanocups Assist the Formation of Ultra-Small Palladium Nanoparticle Catalysts for Hydrogen Evolution Reaction. Materials 2022, 15, 2692. [Google Scholar] [CrossRef]
  55. Quach, Q.; Biehler, E.; Abdel-Fattah, T.M. Synthesis of Copper Nanoparticles Supported over Graphene-like Material Composite as a Catalyst for Hydrogen Evolution. J. Compos. Sci. 2023, 7, 279. [Google Scholar] [CrossRef]
  56. Biehler, E.; Quach, Q.; Abdel-Fattah, T.M. Gold Nanoparticles AuNP Decorated on Fused Graphene-like Materials for Application in a Hydrogen Generation. Materials 2023, 16, 4779. [Google Scholar] [CrossRef]
  57. Biehler, E.; Quach, Q.; Abdel-Fattah, T.M. Synthesis of Platinum Nanoparticles Supported on Fused Nanosized Carbon Spheres Derived from Sustainable Source for Application in a Hydrogen Generation Reaction. Nanomaterials 2023, 13, 1994. [Google Scholar] [CrossRef]
  58. Biehler, E.; Quach, Q.; Abdel-Fattah, T.M. Silver-Nanoparticle-Decorated Fused Carbon Sphere Composite as a Catalyst for Hydrogen Generation. Energies 2023, 16, 5053. [Google Scholar] [CrossRef]
  59. Deraedt, C.; Salmon, L.; Gatard, S.; Ciganda, R.; Hernandez, E.; Ruiz, J.; Astruc, D. Sodium borohydride stabilizes very active gold nanoparticle catalysts. Chem. Commun. 2014, 50, 14194–14196. [Google Scholar] [CrossRef]
  60. Guella, G.; Zanchetta, C.; Patton, B.; Miotello, A. New Insights on the Mechanism of Palladium-Catalyzed Hydrolysis of Sodium Borohydride from 11B NMR Measurements. J. Phys. Chem. B 2006, 110, 17024–17033. [Google Scholar] [CrossRef]
  61. Wen, D.; Liu, W.; Haubold, D.; Zhu, C.; Oschatz, M.; Holzschuh, M.; Wolf, A.; Simon, F.; Kaskel, S. Gold Aerogels: Three-Dimensional Assembly of Nanoparticles and Their Use as Electrocatalytic Interfaces. ACS Nano 2016, 10, 2559–2567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. TEM images of the Pd-fused graphene-like material (PdFGLM) composite material at scale ranges of 20 nm (A), 10 nm (B), 5 nm (C), and 2 nm (D). The white circle showed the layers of FGLM.
Figure 1. TEM images of the Pd-fused graphene-like material (PdFGLM) composite material at scale ranges of 20 nm (A), 10 nm (B), 5 nm (C), and 2 nm (D). The white circle showed the layers of FGLM.
Catalysts 13 01117 g001
Figure 2. EDS spectrum of the PdFGLM material.
Figure 2. EDS spectrum of the PdFGLM material.
Catalysts 13 01117 g002
Figure 3. XRD analysis of the PdFGLM. The asterisk (*) marked the characteristic peaks corresponding to the lattice planes of PdFGLM. The inset shows the P-XRD of FGLM.
Figure 3. XRD analysis of the PdFGLM. The asterisk (*) marked the characteristic peaks corresponding to the lattice planes of PdFGLM. The inset shows the P-XRD of FGLM.
Catalysts 13 01117 g003
Figure 4. FTIR analysis of unsupported FGLM and the PdFGLM composite.
Figure 4. FTIR analysis of unsupported FGLM and the PdFGLM composite.
Catalysts 13 01117 g004
Figure 5. Hydrolysis of NaBH4 as catalyzed by PdFGLM at varying dosages of NaBH4 (625 μmol, 925 μmol, and 1225 μmol).
Figure 5. Hydrolysis of NaBH4 as catalyzed by PdFGLM at varying dosages of NaBH4 (625 μmol, 925 μmol, and 1225 μmol).
Catalysts 13 01117 g005
Figure 6. Hydrolysis of NaBH4 as catalyzed by PdFGLM at varying pH conditions (pH 6, pH 7, and pH 8).
Figure 6. Hydrolysis of NaBH4 as catalyzed by PdFGLM at varying pH conditions (pH 6, pH 7, and pH 8).
Catalysts 13 01117 g006
Figure 7. Hydrolysis of NaBH4 as catalyzed by PdFGLM at varying temperature conditions (283 K, 288 K, 295 K, and 303 K).
Figure 7. Hydrolysis of NaBH4 as catalyzed by PdFGLM at varying temperature conditions (283 K, 288 K, 295 K, and 303 K).
Catalysts 13 01117 g007
Figure 9. Catalytic reusability of PdFGLM after five consecutive trials.
Figure 9. Catalytic reusability of PdFGLM after five consecutive trials.
Catalysts 13 01117 g009
Scheme 1. Assumed mechanism for the hydrolysis of NaBH4 catalyzed by PdFGLM.
Scheme 1. Assumed mechanism for the hydrolysis of NaBH4 catalyzed by PdFGLM.
Catalysts 13 01117 sch001
Table 1. Table of reported activation energies.
Table 1. Table of reported activation energies.
CatalystEa (kJ mol−1)Temperature (K)Reference
Ni71273–308[44]
Raney-Nickel63273–308[44]
Co75273–308[44]
Ru/C67298–358[45]
Ru/Graphite61.1398–318[46]
Pd/C28298–328[47]
Pt–Pd/CNTs19302–332[48]
Au/MWCNTs21.1273–303[49]
Ag/MWCNTs44.5273–303[50]
Pd/MWCNTs62.7273–303[16]
Pt/MWCNTs46.2273–303[51]
BCD-AuNP54.7283–303[52]
PtNPs39.2283–303[53]
PdNPs58.9273–303[54]
CuGLM46.8283–303[55]
AuFGLM45.5283–303[56]
PtFCS53.0283–303[57]
AgNP-FCS37.0283-303[58]
PdFGLM45.1283–303This Work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Quach, Q.; Biehler, E.; Abdel-Fattah, T.M. Synthesis of Palladium Nanoparticles Supported over Fused Graphene-like Material for Hydrogen Evolution Reaction. Catalysts 2023, 13, 1117. https://doi.org/10.3390/catal13071117

AMA Style

Quach Q, Biehler E, Abdel-Fattah TM. Synthesis of Palladium Nanoparticles Supported over Fused Graphene-like Material for Hydrogen Evolution Reaction. Catalysts. 2023; 13(7):1117. https://doi.org/10.3390/catal13071117

Chicago/Turabian Style

Quach, Qui, Erik Biehler, and Tarek M. Abdel-Fattah. 2023. "Synthesis of Palladium Nanoparticles Supported over Fused Graphene-like Material for Hydrogen Evolution Reaction" Catalysts 13, no. 7: 1117. https://doi.org/10.3390/catal13071117

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop